Perovskite: Structural Stability, Oxygen Defect ... - ACS Publications

Feb 22, 2017 - copy (Solartron 1260 Impedance Analyzer). The single crystal was cut into disks coated with a Pt paste on both sides and pressed into t...
0 downloads 0 Views 3MB Size
Article pubs.acs.org/IC

Cubic Sr2ScGaO5 Perovskite: Structural Stability, Oxygen Defect Structure, and Ion Conductivity Explored on Single Crystals Serena Corallini,† Monica Ceretti,*,† Alain Cousson,‡ Clemens Ritter,§ Marco Longhin,† Philippe Papet,† and Werner Paulus† †

Institut Charles Gerhardt, UMR 5253 CNRS-UM-ENSCM, Université Montpellier, Montpellier Cedex 5, France Laboratoire Léon Brillouin, UMR12, CEA-CNRS, CEA Saclay, 91191 Gif Sur Yvette, France § Institut Laue-Langevin, BP 156, 38042 Grenoble, France ‡

S Supporting Information *

ABSTRACT: Oxygen-deficient Sr2ScGaO5 single crystals with a cubic perovskite structure were grown by the floating-zone technique. The transparent crystals of this pure 3D oxygen electrolyte are metastable at ambient temperature, showing one-sixth of all oxygen positions vacant. While neutron single-crystal diffraction, followed by maximum entropy analysis, revealed a strong anharmonic displacements for the oxygen atoms, a predominant formation of ScO6 octahedra and GaO4 tetrahedra is indicated by Raman spectroscopic studies, resulting in a complex oxygen defect structure with short-range order. Temperaturedependent X-ray powder diffraction (XPD) and neutron powder diffraction (NPD) studies reveal the cubic Sr2ScGaO5 to be thermodynamically stable only above 1400 °C, while the stable modification below this temperature shows the brownmillerite framework with orthorhombic symmetry. Cubic Sr2ScGaO5 remains surprisingly kinetically stable upon heating from ambient temperature to 1300 °C, indicating a huge inertia for the retransformation toward the thermodynamically stable brownmillerite phase. Ionic conductivity investigated by impedance spectroscopy was found to be 10−4 S/cm at 600 °C, while oxygen 18O/16O isotope exchange indicates a free oxygen mobility to set in at around 500 °C.



INTRODUCTION Oxygen ion conductors operating at moderate temperatures are materials of major interest for a series of applications in the area of solid-state ionics (solid oxide fuel cells (SOFC), batteries, electrodes, sensors, and catalysts).1,2 Still today high operating temperatures above 800 °C are generally required to obtain reasonable efficiency, limiting their application. In the search for improved oxygen ion conductors, oxygen-deficient perovskite structures play an important role. In this context, oxides with perovskite-type structures have been widely investigated, (Sr,Mg)-doped lanthanum gallate, LaGaO3, being one prominent example as an intermediate-temperature electrolyte with promising performance.3−7 As a consequence of the formation of protonic defects generated by a reaction between oxygen defects and water vapor, oxygen-deficient perovskites are also potential candidates for proton conduction. In this respect oxides with brownmillerite type structures (A2BB′O5) have attracted much attention, especially as they show oxygen ion mobility down to ambient temperature.8−12 This unusual low-temperature oxygen mobility has been evidenced to rely on a phonon-assisted diffusion mechanism, essentially based on large, dynamically activated displacements of the apical oxygen atoms of the BO6 octahedra.13 Oxygen diffusion at room temperature thus makes this class of © XXXX American Chemical Society

compounds attractive for many technologically important applications in the field of solid-state electrolytes and more specifically for membranes and pure electrolytes. Brownmillerite type frameworks containing B-cations with saturated or empty electron shells, such as d0 and d10 configurations, present a special case, as they impose a fixed oxygen stoichiometry, avoiding complications such as structural transitions related to oxygen uptake or release. This makes them good candidates to learn more about the oxygen diffusion mechanisms on a microscopic level, as no changes in the valence state of the B-cations nor their polyhedral arrangement is supposed to occur. In this context, cation-substituted Ba2In2O5 is one of the most studied brownmillerite type electrolytes for SOFC applications.4,14 Recently, Sr 2 ScGaO 5 has been reported as a new brownmillerite phase, showing B-cation ordering of Sc3+ and Ga3+ on octahedral and tetrahedral sites, respectively, adopting the I2mb space-group at ambient temperature.15,16 This symmetry implies an ordered arrangement of the (GaO4)∞tetrahedral chains, which may explain accordingly the absence of oxygen diffusion close to ambient temperature. When NPD Received: December 20, 2016

A

DOI: 10.1021/acs.inorgchem.6b03106 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Empyrean diffractometer (Cu Kα1,2), equipped with an HTK 16N Anton Parr high-temperature chamber. Neutron diffraction on the as-obtained single crystals has been performed on the 5C2 single diffractometer installed at the LLB (ORPHEE reactor in Saclay, France) with λ = 0.83 Å. Such a short wavelength allows having an excellent data quality up to a high momentum transfer of 0.72 Å−1, in order to better explore anisotropic and anharmonic displacement parameters. No absorption correction has been applied due to the negligible absorption coefficients. Complementary T-dependent NPD patterns were obtained on the high-resolution powder diffractometer D2B (λ = 1.59 Å), at the highflux reactor of the Laue-Langevin Institute (ILL, France). Single-crystal structure refinements were carried out using SHELXL,18 while the powder diffraction data were analyzed by the Rietveld method19 using the FullProf suite software.20 Nuclear densities were reconstructed in real space through the maximum entropy method (MEM) via PRIMA (practice iterative MEM analysis);21 scattering lengths were taken from ref 22. Nuclear density distributions were visualized by using the VESTA program.23 The elemental composition and homogeneity were checked by scanning electron microscopy (SEM) analysis using a JEOL JSM 6400 microscope, equipped with an OXFORD INCA EDS instrument for atomic recognition via X-ray fluorescence spectroscopy. SEM/EDS analyses have been performed on a cross section (4 mm in diameter) of the grown crystal, sliced perpendicular to the growth direction. Raman Spectroscopy. Raman spectra were recorded in air by using a LabRAM ARAMIS IR2 (Horiba Jobin Yvon) backscattering micro-Raman spectrometer with helium−neon laser (633 nm) excitation. Impedance Spectroscopy. Ionic conductivity of the as-grown SSGO was measured by alternating current (ac) impedance spectroscopy (Solartron 1260 Impedance Analyzer). The single crystal was cut into disks coated with a Pt paste on both sides and pressed into two Pt disks to ensure bonding. The experiments were carried out at different temperatures (300, 400, 500, and 600 °C) in air at an applied voltage of 0.1 V in the frequency range from 0.1 Hz to 1 MHz. Equivalent circuit modeling was performed by using ZView (Scribner Associates). Isotope Exchange. Polycrystalline SSGO, obtained by crushing a small part of the single crystal, has been characterized by isotopic exchange 18O/16O and thermogravimetric analysis (TGA) coupled with MS. Thermogravimetric measurements were carried out on a NETSCH Jupiter STA 449C thermobalance, equipped with a PFEIFFER VACUUM Thermo Star mass spectrometer.

and synchrotron diffraction were combined together with oxygen isotope 18O/16O exchange experiments, the onset of oxygen mobility was found to occur at 500 °C, which has been interpreted as originating from a subtle phase transition from I2mb toward Imma and from concurrent changes in the lattice dynamics, promoting a dynamic and collective switching of the infinite (GaO4)∞ chains.16 In addition, brownmillerite Sr2ScGaO5 shows a phase transition to the cubic perovskite structure completed at 1500 °C, associated with improved oxygen ion conduction, probably related to the 3D perovskite framework and associated diffusion pathways.17 Since the cubic symmetry can be easily maintained down to ambient temperature via simple furnace cooling, SSGO turns out to be a model system to investigate the influence of the structural dimensionality of the underlying framework for oxygen diffusion mechanisms. We report here on a combination of characterization of the cubic Sr2ScGaO5 single crystal (in the following indicated as cSSGO). All the studies reported so far on such material were performed on polycrystalline samples. Our effort to dispose large and high-quality crystals of Sr2ScGaO5 goes along with our motivation to study anisotropic oxygen mobility together with accurate structural analysis and lattice dynamics with inelastic neutron scattering. It is also a rare case where single crystals can be obtained for cubic and stoichiometric oxygendeficient perovskite, suitable for high-resolution X-ray and neutron diffraction studies. High-resolution structure analysis has been performed using X-ray and neutron diffraction methods. Single-crystal neutron diffraction with subsequent analysis of the nuclear scattering density by the maximum entropy method has been performed in order to describe in more detail oxygen displacement factors and associated diffusion pathways. To better understand the oxygen mobility mechanisms, these studies were complemented by Raman and impedance spectroscopy as well as by the determination of the oxygen isotope exchange behavior via thermogravimetric analysis coupled mass spectroscopy studies on 18O-enriched cSSG18O, under a 16O2 atmosphere.





RESULTS AND DISCUSSION When polycrystalline Sr2ScGaO5 is synthesized at 1200 °C by a classical solid-state reaction, the brownmillerite modification is obtained, adopting at ambient temperature the I2mb space group with unit cell parameters a = 5.91048(5) Å, b = 15.1594(1) Å, and c = 5.70926(4) Å.16 Heating to 1600 °C leads, however, to the formation of the cubic perovskite structure, which is easily preserved when the sample is cooled to ambient temperature.17 The cubic perovskite phase is also obtained under single-crystal growth conditions using the floating-zone method reported above. The chemical composition of the sample and its homogeneity were verified on a section cut from the as-grown c-SSGO crystal. It was found that the atomic percentages of scandium, gallium, and strontium remain constant over the whole cross section, while the Ga/Sc ratio was found to be 1 throughout the whole cross-section (see Figure S3 in the Supporting Information), in agreement with the chemical formula of Sr2ScGaO5. The phase purity of the grown single crystals was checked by X-ray powder diffraction on small sections of the asgrown single crystals that were crushed into powder and used for all powder diffraction experiments reported below. XRD patterns of cubic SSGO (Figure 1) could be indexed with a

EXPERIMENTAL SECTION

Single-Crystal Growth. Polycrystalline Sr2ScGaO5 with a brownmillerite structure was prepared using standard solid-state chemistry synthesis at high temperature. High-purity SrCO 3 (99,995%, Aldrich), Ga2O3 (99,99%, Aldrich), and Sc2O3 (REacton 99.99% REO, Alfa Aesar) were thoroughly mixed and heated in air at 1200 °C for 48 h. Seed and feed rods for crystal growth were obtained by hydrostatically pressing the obtained Sr2ScGaO5 powder at 10 bar in a cylindrical latex tube of 6 mm in diameter and 100 mm in length. The as-obtained rods were then sintered in air at 1200 °C for 24 h to obtain dense polycrystalline rods. Large, high-quality c-SSGO single crystals have been obtained and characterized as reported elsewhere.17 Single-crystal growth was carried out by the floating-zone method using a mirror furnace (NEC SC2, Japan) equipped with two 500 W halogen lamps as heat sources and two ellipsoidal mirrors. Crystal growth was performed in air with a typical traveling rate of 2 mm/h, while the upper and lower shafts were rotated in opposite directions at 25 rpm. Structural Analysis. Several characterizations were performed on the as-grown Sr2ScGaO5 single crystals. Phase purity and structure were checked by laboratory X-ray powder diffraction on a Bruker AXS D8 Advance diffractometer using Cu Kα1 radiation. T-dependent Xray powder patterns up to 1600 °C were collected on a Panalytical B

DOI: 10.1021/acs.inorgchem.6b03106 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

GaO4 tetrahedra beside ScO6 octahedra. This could then be equivalent to the formation of small differently oriented microdomains of the brownmillerite phase, as already reported for polycrystalline Sr2Ga1.5Sc0.5O5 and Sr2Co2−xAlxO5.15,25 Such a scenario can be expected in view of the rigid character and stability of both polytypes of polyhedra and would imply consequences for related oxygen diffusion pathways and energies. Figure 2 depicts powder X-ray diffraction patterns from room temperature up to 1600 °C, obtained during a heating cycle.

cubic perovskite cell. No extra reflections are detectable, showing the high purity of the single crystal.

Figure 1. Rietveld refinement on the X-ray powder diffraction data (Cu Kα1) at room temperature of the ground single crystal (RBragg = 4.44%, χ2 = 3.16, Rp = 8.39%, Rpw = 11.7%).

Structure Analysis by Neutron and X-ray Diffraction. Structure refinement has been done by a combination of X-ray and neutron diffraction experiments. In particular, considering the difference between the neutron scattering lengths b of scandium (12.29 fm) and gallium (7.29 fm), neutron diffraction data were used to determine the overall B-cation and oxygen occupancies (b = 5.80 fm). Structure refinements from NPD data, using the Fullprof software,24 yielded for polycrystalline c-SSGO an oxygen vacancy perovskite structure with space group Pm3̅m and a = 3.9932(1) Å, corresponding to a unit cell volume of 63.6736(9) Å3 (Figure 1 and Table 1). In contrast to orthorhombic Sr2ScGaO5, with gallium exclusively on tetrahedral sites and scandium on octahedral sites, c-SSGO implies, from a purely crystallographic point of view, an average Sc/Ga distribution on the B-site and accordingly a statistical distribution of the 1/6 oxygen vacancies. However, despite this average structural description, one might not exclude the existence of possible short-range ordering, especially related to the formation of

Figure 2. X-ray diffraction patterns of c-SSGO as a function of the temperature. On the right is shown an enlargement of a particular 2θ region where the formation of the brownmillerite phase is evident at 1400 °C. The very intense reflection marked with an asterisk at 2θ = 46.2° originates from the Pt sample holder.

Once obtained at ambient temperature, cubic c-SSGO shows a remarkable stability up to high temperatures. Orthorhombic Sr2ScGaO5 is obtained again only above 1400 °C, while we noticed that a quantitative cubic to orthorhombic transformation proceeds on a time scale of only several hours.

Table 1. Structural Parameters after Rietveld Refinements of NPD Data Obtained for Crushed c-SSGO Crystal at Different Temperaturesa a = b = c (Å) V (Å3) RBragg (%) Rwp (%) Rp (%) χ2 Sr Uiso (Å2) Sc Uiso (Å2) Ga Uiso (Å2) O U11 (Å2) U22 = U33 (Å2) Ueq (Å2) O occ

T = 25 °C

T = 400 °C

T = 500 °C

T = 800 °C

T = 1000 °C

3.9932(1) 63.6774(9) 4.6 6.6 4.9 2.8 0.0267 (7) 0.0280(6) 0.0280(6) 0.0398(2) 0.092(2) 0.0772(18) 2.458(26)

4.0101(1) 64.4877(10) 3.1 6.1 4.8 2.1 0.0361(8) 0.0323(6) 0.0323(6) 0.042(2) 0.097(2) 0.0823(19) 2.446(26)

4.0148(1) 64.7144(10) 5.2 6.3 4.8 2.1 0.0359(8) 0.0338(6) 0.0338(6) 0.050(2) 0.0945(18) 0.0797(19) 2.433(27)

4.0303(1) 65.4661(13) 4.1 6.4 4.8 1.9 0.0444(9) 0.0397(7) 0.0397(7) 0.059(2) 0.104(2) 0.089(2) 2.421(29)

4.0412(1) 65.9995(12) 4.9 5.9 4.4 2.7 0.0490(10) 0.0437(8) 0.0437(8) 0.059(3) 0.113(3) 0.095(3) 2.431(27)

a Diffraction data have been obtained on the high-resolution powder diffractometer D2B at the ILL (λ = 1.59 Å). Atomic positions are the same as those reported in Table 2.

C

DOI: 10.1021/acs.inorgchem.6b03106 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry The pure cubic phase forms again on further heating to 1600 °C. Following these order−disorder transitions of SSGO, it becomes obvious that the cubic perovskite structure is kinetically stabilized during cooling but is the thermodynamically stable phase at 1600 °C. At ambient temperature, the cubic phase is thus metastable, while the orthorhombic SSGO is the thermodynamically stable phase. The kinetics of the observed phase transformations is somehow contrary to what is observed for other brownmillerite phases. Sr(Fe/Co)O2.5 shows, for example, a very rapid transformation from the oxygen-deficient cubic perovskite phase at high temperature toward the brownmillerite framework, even when it is quenched in liquid N2.11,26 While Fe and Co can easily adapt to rapid changes in the coordination from octahedral to pyramidal and tetrahedral, no such correspondence exists either for GaO4 tetrahedra or for ScO6 octahedra having a more “rigid” character and also showing a “fixed” trivalent oxidation state for both spherical 3d metal atoms. We studied the structure of c-SSGO by single-crystal neutron diffraction at room temperature, while X-ray and neutron powder diffraction studies have been done as a function of the temperature. Again, the polycrystalline samples have been obtained by crushing a c-SSGO single crystal. Results of the Tdependent structure refinements of the c-SSGO phase up to 1000 °C, obtained from NPD by the Rietveld method, are presented in Table 1 (the corresponding diffraction patterns are shown in Figure S4 in the Supporting Information). For all refinements, the Sr occupancy has been fixed to unity, while oxygen site occupancies were refined. The B-cation site occupancy was constrained to be unity, inversely correlating Ga and Sc contributions. The deviation from an equal Sc/Ga occupation on the B-cation site was found to be less than 2%; the same error bar holds for the oxygen occupancies, which are statistically vacant by 1/6. Strongly anisotropic oxygen displacement factors become evident, giving rise to a disk-shaped distribution of all oxygen atoms, as shown in Figure 4 (top). The U22 (U33) values (Figure 3b) do not converge toward 0 with decreasing temperature, indicating that some significant static displacements still persist at low temperatures. We underline a very good agreement between the ADP obtained by single-crystal (Table 2) and powder data (Table 1), excluding systematic errors, to be at the origin for the high U22 = U33 values found for the oxygen atoms. This is further supported by the fact that, for both Sr and Sc/Ga atoms, a continuous decrease in their respective isotropic displacement factors is found (see Table 1) that is almost identical with the U11 parameters of the oxygen atoms (Figure 3b). In order to get a better insight into especially the oxygen displacements, we investigated the respective scattering densities by high-resolution single-crystal neutron diffraction up to high momentum transfers and subsequent analysis of the nuclear densities by maximum entropy reconstruction. In this approach, the only structural constraint is imposed by the space group, and thus the resulting displacements are equivalent to an anharmonic description of the scattering densities. Values from a classical structure refinement obtained with SHELXL are similar to what has been obtained from NPD at ambient temperature (see Table 2). The scattering densities obtained from maximum entropy analysis (Figure 4, bottom) reveal, however, a more complex behavior of the oxygen displacements, becoming cylindrically shaped with complex deepening. This directly reflects the possible presence of local distortions

Figure 3. (a) Variation of the cubic lattice parameter a as a function of the temperature for c-SSGO. The red point presents the “average” brownmillerite lattice parameter, reduced to the cubic phase. (b) Evolution of the displacement parameter Uij, as a function of the temperature. The oxygen anisotropic displacements U11 and U22 (=U33) are represented in green, while the isotropic Uiso points for Sc/Ga and Sr atoms are shown in blue and orange, respectively. Open red circles at room temperature refer to anisotropic oxygen displacement factors, determined by single-crystal neutron diffraction, indicating a very good agreement between single-crystal and powder data. All error bars are smaller than the symbol size.

Table 2. Sr2ScGaO5 Structure Data Obtained from SingleCrystal Neutron Diffraction Data, Collected on Diffractometer 5C2@LLB with λ = 0.83 Å, Installed at the Hot Source of the ORPHEE Reactora atom

x

y

z

occ

U (Å2)

Sr Sc Ga O

0 1/2 1/2 0

0 1/2 1/2 1/2

0 1/2 1/2 1/2

1.00 0.521(41) 0.479(41) 2.462(42)

Uiso = 0.0310(17) Uiso = 0.0312(13) Uiso = 0.0312(13) U11 = 0.0390(32) U22 = U33 = 0.0901(48)

a Conditions and details: refinements carried out in space group Pm3̅m with unit cell parameter a = 3.9932 Å (determined from XRPD, Cu Kα1), 152 measured reflections (sin θ/λ = 0.72 Å−1) (49 unique), Rint(hkl) = 3.1%, Rw = 10%, GOF = 1.23.

when considering GaO4 tetrahedra and respective ScO6 octahedra. The values for both Sr and Sc/Ga displacements remain, however, comparable to the Debye−Waller factors found from the classical refinements given in Tables 1 and 2. Interestingly the anisotropic displacement factors found for cD

DOI: 10.1021/acs.inorgchem.6b03106 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Figure 5. Raman spectrum of cubic oxygen-deficient Sr2ScGaO5 with a perovskite structure, showing two separate peaks around 600 and 750 cm−1, associated with the stretching modes of ScO6 octahedra and GaO4 tetrahedra, respectively.

consequence of the oxygen vacancies and related deviations from a local cubic symmetry.29 Similar features have also been found in the proton-conducting cubic perovskite BaInxZr1−xO3−x/2.30 The two well-distinguished peaks correspond to the (Sc/Ga)−O stretching modes30,31 and reinforce the hypothesis that in a first approach two distinct polyhedra, i.e. GaO4 tetrahedra and ScO6 octahedra, are present in cSSGO. Impedance Spectroscopy. In order to directly access oxygen ion mobility, ac impedance measurements have been performed on the as grown single crystals up to 600 °C along the [110] direction, corresponding to the crystal growth direction, as identified by Laue backscattering (see Figure S2 in the Supporting Information). As expected from a single-crystal study, the complex impedance spectrum of c-SSGO consists of single broad semicircles (Figure 6) for all temperatures. Because the data were obtained on a single crystal, with a semicircle passing through the origin, they can be attributed to a bulk phenomenon and not to grain boundaries. The associated

Figure 4. (Sc/Ga)Ox polyhedra in c-SSGO, obtained at room temperature from neutron single-crystal diffraction as a harmonic description of the displacement factors (top) and the reconstructed nuclear scattering density using the maximum entropy method (bottom). The perovskite unit cell and the (Sc/Ga)O6 octahedra are outlined. While isotropic displacements are found for Sr and (Sc/Ga), the isosurfaces of oxygen atoms reveal an anisotropic cylindrical shape with significant deepening.

SSGO are almost identical with what has been found for the homologous high-temperature perovskite Ba2In2O5 at 980 °C.27 Such strong displacements could therefore be expected to be more generally present in oxygen-deficient perovskite and have to be taken into account, especially when the distribution of protonic defects around oxygen positions in similar types of compounds is described. For the proton-conducting perovskite BaZr1/2In1/2O2.5(OD)1/2 where, however, all oxygen sites are completely filled, still one-third of the displacement factors found for the high-temperature modification of Ba2In2O5 have been reported.28 For c-SSGO, the large displacements are supposed to result from distortions of the different (Ga/Sc)Ox polyhedra, since a strongly developed lattice dynamics, inducing a continuous and dynamically activated reorganization of the (Ga/Sc)Ox polyhedra and related oxygen defects, is unlikely to be present at ambient temperature. Raman Spectroscopy. For an ideal cubic perovskite with the Pm3̅m space group, no first-order Raman-active vibration frequencies are allowed. As shown in Figure 5, two relatively intense but broad peaks, centered at 600 and 743 cm−1, are nevertheless observed for c-SSGO. Their appearance is a

Figure 6. Cole−Cole plot obtained from ac impedance measurements on a c-SSGO single crystal at different temperatures. The inset shows an enlargement for the values obtained at higher temperatures. E

DOI: 10.1021/acs.inorgchem.6b03106 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry capacitance of about 6 pF has been calculated using the usual relationship ωRC = 1 at the top of each arc for all investigated temperatures.32 The bulk resistivity was modeled with an equivalent circuit made of a resistor (R) in parallel with a constant phase element (CPE). A second resistor (RΩ) was added in series to model ohmic losses originating from the setup. Figure 7 shows the Arrhenius plot of the bulk

Figure 8. TGA-MS spectroscopic study on Sr2ScGa18O5, enriched to above 95% of 18O. The experiment was performed under a 16O2 (20%)/He (80%) atmosphere with a heating rate of 5 K. The decrease in mass (black line) observed is 2.65% and sets in above 500 °C. The red and green lines correspond to the intensities of the 18O−18O and 18 O−16O mass signals, respectively.

above 500 °C, that correlates with the 18O/16O isotope exchange. Due to slow exchange kinetics even at quite elevated temperatures, complete exchange is only observed at the end of the plateau at 850 °C, corresponding to almost the theoretical mass loss of 2.60%. This result is quite similar to what had been obtained on polycrystalline o-SSGO16 with a brownmilleritetype structure. These results imply on an absolute temperature scale a fairly low onset temperature for free oxygen diffusion, especially when SSGO is considered as a pure oxygen electrolyte. The relatively low exchange kinetics indicated by 18 O/16O isotope exchange underpins again the hypothesis that on a local level the cubic perovskite SSGO phase might consist of brownmillerite-type microdomains with 1D oxygen diffusion pathways, even though it appears overall to be cubic, as shown from diffraction studies.

Figure 7. Temperature-dependent Arrhenius plot of the bulk conductivity as obtained from the ac impedance measurements on the c-SSGO single crystal.

conductivity data as resulting from the ac impedance measurements reported in Figure 6. The values are similar to those reported for the oxygen ionic conductor brownmillerite Ba2In2O514 and for its cubic perovskite modification33 in the corresponding temperature range. The activation energy for a c-SSGO single crystal of 0.67 eV has been determined by a linear regression of the conductivity data. This value is, however, significantly lower in comparison to the 1.34 eV obtained for polycrystalline Sr2Ga1.5Sc0.5O5,15 showing the same oxygen-deficient cubic perovskite structure. This important difference is not supposed to be related to the change of the Bsite cation stoichiometry, and it cannot even be attributed to any proton mobility, as the absence of water was previously confirmed for the single crystal. Hence, we attribute the lower activation energy of c-SSGO to the absence of grain boundaries in comparison to the polycrystalline SSGO sample reported in ref 15. The availability of single-crystalline electrodes for moderate temperature oxygen ion conductivity thus becomes important especially for this lower temperature range. Oxygen 18O/16O Isotope Exchange Studies. In order to investigate the temperature at which oxygen mobility appears, we performed TG experiments on polycrystalline 18O-enriched c-SSGO. Experiments were carried out under an atmosphere consisting of O2 (20%)/He (80%), equivalent to the oxygen partial pressure in air. The onset temperature for “free” oxygen mobility is then given by the weight change associated with the 18 O vs 16O exchange. For a 100% enriched 18O sample the corresponding mass change theoretically yields 2.65%. Figure 8 shows the result of a TGA coupled MS study of cSSG18O, enriched with 18O fairly close to 100%, obtained with a heating rate of 5 K/min up to 850 °C. The temperature was held for 3 h at 850 °C, before it was lowered to ambient temperature again at 5 K/min. The detection of 18O−16O and pure 18O2 molecules was simultaneously analyzed by mass spectroscopy. The TG signal reveals a mass reduction, starting



CONCLUSIONS The structure and stability range of cubic Sr2ScGaO5 has been investigated by neutron and X-ray diffraction studies. While the compound forms only above 1400 °C, it remains kinetically stabilized during furnace cooling down to ambient temperature. This allows the growth of high-quality large single crystals of cSr2ScGaO5 with an oxygen-deficient perovskite structure, suitable for neutron structure investigations. The oxygen displacement factors reveal strong anisotropic cylindricalshape behavior, as determined by maximum entropy analysis. This suggests significant deviations of the oxygen atoms from their equilibrium positions in the perovskite framework. For oxygen-deficient perovskites containing protonic defects, the proton positions should consequently be considered to be localized significantly far away from the ideal oxygen positions, which is interesting as an input parameter for simulations. From Raman spectroscopic studies the presence of two distinct peaks in the typical regions of M−O stretching modes supports the formation of well-defined GaO4 and ScO6 polyhedra and consequently the probable formation of brownmillerite-type microdomains within the cubic symmetry. It thus becomes clear that Sr2ScGaO5 is not comparable by far to the idealized perovskite framework but shows an anionic sublattice with a complex distortion of the oxygen positions. Importantly, we could show that the activation energy of 0.67 eV for the cubic F

DOI: 10.1021/acs.inorgchem.6b03106 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

(9) Ichikawa, N.; Iwanowska, M.; Kawai, M.; Calers, C.; Paulus, W.; Shimakawa, Y. Reduction and oxidation of SrCoO2.5 thin films at low temperatures. Dalton Trans. 2012, 41, 10507−10510. (10) Jeen, H.; Choi, W. S.; Biegalski, M. D.; Folkman, C. M.; Tung, I. C.; Fong, D. D.; Freeland, J. W.; Shin, D.; Ohta, H.; Chisholm, M. F.; Lee, H. N. Reversible redox reactions in an epitaxially stabilized SrCoOx oxygen sponge. Nat. Mater. 2013, 12, 1057−1063. (11) Le Toquin, R.; Paulus, W.; Cousson, A.; Prestipino, C.; Lamberti, C. Time-resolved in situ studies of oxygen intercalation into SrCoO2.5, performed by neutron diffraction and X-ray absorption spectroscopy. J. Am. Chem. Soc. 2006, 128, 13161−13174. (12) Nemudry, A.; Weiss, P.; Gainutdinov, I.; Boldyrev, V.; Schollhorn, R. Room temperature electrochemical redox reactions of the defect perovskite SrFeO2.5+x. Chem. Mater. 1998, 10, 2403−2411. (13) Paulus, W.; Schober, H.; Eibl, S.; Johnson, M.; Berthier, T.; Hernandez, O.; Ceretti, M.; Plazanet, M.; Conder, K.; Lamberti, C. Lattice Dynamics To Trigger Low Temperature Oxygen Mobility in Solid Oxide Ion Conductors. J. Am. Chem. Soc. 2008, 130, 16080− 16085. (14) Goodenough, J. B.; Ruiz-Diaz, J. E.; Zhen, Y. S. Oxide-ion conduction in Ba2In2O5 and Ba3In2MO8 (M = Ce, Hf, or Zr). Solid State Ionics 1990, 44, 21−31. (15) Chernov, S. V.; Dobrovolsky, Y. A.; Istomin, S. Y.; Antipov, E. V.; Grins, J.; Svensson, G.; Tarakina, N. V.; Abakumov, A. M.; Van Tendeloo, G.; Eriksson, S. G.; Rahman, S. M. H. Sr2GaScO5, Sr10Ga6Sc4O25, and SrGa0.75Sc0.25O2.5: a Play in the Octahedra to Tetrahedra Ratio in Oxygen-Deficient Perovskites. Inorg. Chem. 2012, 51, 1094−1103. (16) Corallini, S.; Ceretti, M.; Silly, G.; Piovano, A.; Singh, S.; Stern, J.; Ritter, C.; Ren, J.; Eckert, H.; Conder, K.; Chen, W.-t.; Chou, F.-C.; Ichikawa, N.; Shimakawa, Y.; Paulus, W. One-Dimensional Oxygen Diffusion Mechanism in Sr2ScGaO5 Electrolyte Explored by Neutron and Synchrotron Diffraction, 17O NMR, and Density Functional Theory Calculations. J. Phys. Chem. C 2015, 119, 11447−11458. (17) Ceretti, M.; Corallini, S.; Paulus, W. Influence of Phase Transformations on Crystal Growth of Stoichiometric Brownmillerite Oxides: Sr2ScGaO5 and Ca2Fe2O5. Crystals 2016, 6, 146. (18) Sheldrick, G. A short history of SHELX. Acta Crystallogr., Sect. A: Found. Crystallogr. 2008, 64, 112−122. (19) Rietveld, H. M. Line Profile of neutron powder diffraction peaks for structure refinement. Acta Crystallogr. 1967, 22, 151−155. (20) Frontera, C.; Rodriguez-Carvajal, J. FULLPROF as a new tool for flipping ratio analysis. Phys. B 2003, 335, 219−222. (21) Izumi, F.; Dilanian, R. A. In Recent Research Developments in Physics; Pandalai, S. G., Ed.; Transworld Research Network, Trivandrum: 2002; Vol. 3, pp 699−726. (22) Koester, L. In Neutron Physics; Springer: Berlin, Heidelberg, 1977; Vol. 80, Chapter 1, pp 1−55. (23) Momma, K.; Izumi, F. VESTA 3 for three-dimensional visualization of crystal, volumetric and morphology data. J. Appl. Crystallogr. 2011, 44, 1272−1276. (24) Rodríguez-Carvajal, J. Recent advances in magnetic structure determination by neutron powder diffraction. Phys. B 1993, 192, 55− 69. (25) Lindberg, F.; Svensson, G.; Istomin, S. Y.; Aleshinskaya, S. V.; Antipov, E. V. Synthesis and structural studies of Sr2Co2−xAlxO5, 0.3 ⩽ x ⩽ 0.5. J. Solid State Chem. 2004, 177, 1592−1597. (26) Gallagher, P. K.; MacChesney, J. B.; Buchanan, D. N. E. Mössbauer Effect in the System SrFeO2.5−3.0. J. Chem. Phys. 1964, 41, 2429−2434. (27) Berastegui, P.; Hull, S.; García-García, F. J.; Eriksson, S. G. J. Solid State Chem. 2002, 164, 119. (28) Ahmed, I.; Knee, C. S.; Karlsson, M.; Eriksson, S. G.; Henry, P. F.; Matic, A.; Engberg, D.; Börjesson, L. J. Alloys Compd. 2008, 450, 103. (29) Siny, I. G.; Katiyar, R. S.; Bhalla, A. S. Cation arrangement in the complex perovskites and vibrational spectra. J. Raman Spectrosc. 1998, 29, 385−390.

Sr2ScGaO5 single crystal is only half the value reported for polycrystalline Sr2Ga1.5Sc0.5O5, showing the same oxygen defect perovskite structure. This becomes specifically important at lower temperatures and underlines the interest in singlecrystalline electrodes, e.g. in the form of epitaxial thin films, for oxygen ion conductivity at moderate temperatures.



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.inorgchem.6b03106. Crystal growth details, single-crystal characterization data (neutron rocking curves, chemical composition, Laue pattern), and neutron powder diffraction patterns as a function of the temperature (PDF)



AUTHOR INFORMATION

Corresponding Author

*E-mail for M.C.: [email protected]. ORCID

Monica Ceretti: 0000-0001-9704-8251 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This work was partially financially supported by the French National Research Agency, through the project AMOXIS (No. ANR-14-CE05-0016-02). We are grateful to Dr. B. Fraisse and Dr. D. Bourgogne of the “Plateforme d’Analyse et de Caractérisation du Pôle Chimie Balard”, Montpellier, France, for X-ray thermodiffraction and Raman measurements. We thank the Laboratoire Léon Brillouin (LLB, Saclay, France) and the Institut Laue Langevin (ILL Grenoble, France) for the allocation of neutron beam time. We acknowledge Prof. K. Conder, PSI-Villigen (Villigen, Switzerland), for TGA measurements.



REFERENCES

(1) Skinner, S. J.; Kilner, J. A. Oxygen ion conductors. Mater. Today 2003, 6, 30−37. (2) Marques, F. M. B.; Kharton, V. V.; Naumovich, E. N.; Shaula, A. L.; Kovalevsky, A. V.; Yaremchenko, A. A. Oxygen ion conductors for fuel cells and membranes: selected developments. Solid State Ionics 2006, 177, 1697−1703. (3) Ishihara, T.; Matsuda, H.; Takita, Y. Doped LaGaO3 Perovskite Type Oxide as a New Oxide Ionic Conductor. J. Am. Chem. Soc. 1994, 116, 3801−3803. (4) Goodenough, J. B.; Manthiram, A.; Kuo, J. F. Oxygen diffusion in perovskite-related oxides. Mater. Chem. Phys. 1993, 35, 221−224. (5) Steele, B. C. H.; Heinzel, A. Materials for fuel-cell technologies. Nature 2001, 414, 345−352. (6) Tao, S. W.; Irvine, J. T. S.; Kilner, J. A. An Efficient Solid Oxide Fuel Cell Based upon Single-Phase Perovskites. Adv. Mater. 2005, 17, 1734−1737. (7) Jacobson, A. J. Materials for Solid Oxide Fuel Cells. Chem. Mater. 2010, 22, 660−674. (8) Hibino, M.; Harimoto, R.; Ogasawara, Y.; Kido, R.; Sugahara, A.; Kudo, T.; Tochigo, E.; Shibata, N.; Ikuhara, Y.; Mizuno, N. A New Rechargeable Sodium Battery Utilizing Reversible Topotactic Oxygen Extraction/Insertion of CaFeOz (2.5 ≤ z ≤ 3) in an Organic Electrolyte. J. Am. Chem. Soc. 2014, 136, 488−494. G

DOI: 10.1021/acs.inorgchem.6b03106 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry (30) Karlsson, M.; Matic, A.; Knee, C. S.; Ahmed, I.; Eriksson, S. G.; Börjesson, L. Short-Range Structure of Proton-Conducting Perovskite BaInxZr1‑xO3‑x/2 (x = 0−0.75). Chem. Mater. 2008, 20, 3480−3486. (31) Bielecki, J.; Parker, S. F.; Ekanayake, D.; Rahman, S. M. H.; Borjesson, L.; Karlsson, M. Short-range structure of the brownmillerite-type oxide Ba2In2O5 and its hydrated proton-conducting form BaInO3H. J. Mater. Chem. A 2014, 2, 16915−16924. (32) Bauerle, J. E. Study of solid electrolyte polarization by a complex admittance method. J. Phys. Chem. Solids 1969, 30, 2657−2670. (33) Yamamura, H.; Hamazati, H.; Kakinumai, K.; Moori, T.; Haneda, H. Order-disorder transition and Electrical Conductivity of the brownmillerite solid solutions System Ba2(In,M)2O5 (M = Ga,Al). J. Korean Phys. Soc. 1999, 35, S200−S204.

H

DOI: 10.1021/acs.inorgchem.6b03106 Inorg. Chem. XXXX, XXX, XXX−XXX